Tintinalli's Emergency Medicine: A Comprehensive Study Guide, 9e

Chapter 12: Approach to Nontraumatic

Bret A. Nicks; John P. Gaillard

FIGURE 12-1.

EPIDEMIOLOGY

Using a systolic <90 mm Hg as criteria, 0.4% to 1.3% of patients presenting to EDs are in shock.1 Mortality depends on the inciting event. has an estimated hospital mortality of 26%.2 has an estimated hospital mortality of 39% to 48%.3,4 occurs in <20% of spinal cord injuries (cervical, 19.3%; thoracic, 7%; lumbar, 3%).5 The definition of and treatment approach to shock continue to evolve, but the initial approach to a patient in shock follows similar principles, regardless of the inciting factors or cause.

Patients present to the ED in varying stages of critical illness and shock. These stages are confounded by age, comorbidities, and delays in presentation. A focus on early recognition, rapid diagnosis, and empiric is essential. Therapy and patient stabilization may need to occur simultaneously with evaluation.

PATHOPHYSIOLOGY

Shock is a state of circulatory insuiciency that creates an imbalance between tissue oxygen supply (delivery) and demand (consumption), resulting in end-organ dysfunction. Reduction in eective perfusion may be due to a local or global delivery deficiency or utilization deficiency with suboptimal substrate at the cellular or subcellular level.6-8 The mechanisms that can result in shock are frequently divided into four categories: (1) hypovolemic, (2) distributive, (3) cardiogenic, and (4) obstructive.

CATEGORIES OF SHOCK

The four categories of shock can be described in terms of their respective physiologic changes and common causes, recognizing that a single etiology may manifest the clinical findings of more than one shock type (Table 12-1).9,10 occurs when decreased intravascular fluid or decreased blood volume causes decreased preload, stroke volume, and cardiac output (CO). Severe blood loss (hemorrhage) can cause decreased myocardial oxygenation, which decreases contractility and CO. This action may lead to an autonomic increase in the systemic vascular resistance (SVR). Hypovolemic shock can also occur due to volume loss from other etiologies. In , there is relative intravascular volume depletion due to marked systemic vasodilatation. This is most commonly seen in septic shock.6 Compensatory responses to decreased SVR may include increased CO (increased contractility and heart rate) and . The concurrent decreased SVR results in a decreased preload and may hinder CO overall. In sepsis, up to 40% of patients may have a transient cardiomyopathy characterized by decreased contractility and increased mortality.6 , adrenal insuiciency, and neurogenic shock are additional causes of distributive shock. In cardiogenic shock, the le ventricle fails to deliver oxygenated blood to peripheral tissues due to variances in contractility, as well as preload, aerload, and right ventricular function.8 Myocardial infarction is the most common cause of cardiogenic shock. Dysrhythmias are another common cause because they can lead to a decreased CO. Bradyarrhythmias result in low CO, and tachyarrhythmias can result in decreased preload and stroke volume. Patients with cardiogenic shock may soon develop clinically evident infection (up to 46%) and/or demonstrate an inflammatory response similar to but less pronounced than with septic shock.9 is uncommon (1%) and is due to a decrease in venous return or cardiac compliance due to an increased Loading [Contrib]/a11y/accessibility-menu.js le ventricular outflow obstruction or marked preload decrease. Cardiac tamponade, pulmonary embolism, and tension pneumothorax are causes of obstructive shock.

TABLE 12-1 Categories of Shock

Type Percentage* Hemodynamic Changes Etiologies†

Distributive 33%–50% Decreased preload, decreased SVR, mixed CO Sepsis, neurogenic shock, anaphylaxis

Hypovolemic 31%–36% Decreased preload, increased SVR, decreased CO Hemorrhage, capillary leak, GI losses, burns

Cardiogenic 14%–29% Increased preload, increased aerload, increased SVR, MI, dysrhythmias, heart failure, valvular decreased CO disease

Obstructive 1% Decreased preload, increased SVR, decreased CO PE, pericardial tamponade, tension PTX

*Percentage of patients presenting to the ED; percentages will vary depending on hospital type and population.

†Etiologies may demonstrate findings of more than one type of shock.

Abbreviations: CO = cardiac output; MI = myocardial infarction; PE = pulmonary embolism; PTX = pneumothorax; SVR = systemic vascular resistance.

FACTORS AFFECTING CARDIAC OUTPUT

CO is determined by heart rate and stroke volume. Stroke volume is dependent on preload, aerload, and contractility. The mean arterial pressure is dependent on CO and the SVR. This is important because there is a mean arterial pressure threshold below which oxygen delivery is decreased. SVR directly impacts mean arterial pressure, but also impacts aerload and thus CO. Patients in shock may initially have normal blood pressures (cryptic shock), yet have other objective signs of shock (see later section “Clinical Features”). Tissue oxygenation is predicated on CO being suicient enough to deliver oxygenated hemoglobin to the tissues. CO is dependent on the interplay of cardiac inotropy (speed and shortening capacity of myocardium), chronotropy (heart contraction rate), and lusitropy (ability of heart muscle to relax and heart chambers to fill). Determinants of inotropy include autonomic input from sympathetic activation, parasympathetic inhibition, circulating catecholamines, and short-lived responses to an increase in aerload (Anrep eect) or heart rate (Bowditch eect). Increases in the inotropic state help to maintain stroke volume at high heart rates.11 During shock states, higher levels of epinephrine will be produced and reinforce adrenergic tone. Epinephrine levels are significantly elevated during induced hemorrhagic shock, but these levels subsequently reduce to almost normal levels aer adequate blood pressure is restored.12 An acidotic milieu, which is common in shock, further compromises ventricular contractile force and blood pressure.13 Chronotropy and lusitropy are both influenced by sympathetic input. Norepinephrine interacts with cardiac β1-receptors, resulting in increased cyclic adenosine monophosphate. This leads to a process of intracellular signaling with an increased chronotropy and sequestration of calcium, leading to myocardial relaxation.11

LACTIC ACID

When compensatory mechanisms fail to correct the imbalance between tissue supply and demand, anaerobic metabolism occurs and results in the formation of lactic acid. Lactic acid is rapidly buered, resulting in the formation of measured serum lactate. NoLromadainl gv [eCnoonturisb ]l/aac1t1ay/taec clesvseiblsil iaty-rem

COMPENSATORY MECHANISMS AND THEIR FAILURE

Shock provokes a myriad of autonomic responses to maintain perfusion pressure to vital organs. Stimulation of the carotid baroreceptor stretch reflex activates the sympathetic nervous system, triggering (1) arteriolar vasoconstriction, resulting in redistribution of blood flow from the skin, skeletal muscle, kidneys, and splanchnic viscera; (2) an increase in heart rate and contractility that increases CO; (3) constriction of venous capacitance vessels, which augments venous return; (4) release of the vasoactive hormones epinephrine, norepinephrine, dopamine, and cortisol to increase arteriolar and venous tone; and (5) release of antidiuretic hormone and activation of the renin-angiotensin axis to enhance water and sodium conservation to maintain intravascular volume.13

These compensatory mechanisms attempt to maintain oxygen delivery to the most critical organs (heart and brain), but blood flow to other organs, such as the kidneys and GI tract, may be compromised. The cellular response to decreased oxygen delivery (adenosine triphosphate depletion) leads to ion-pump dysfunction, influx of sodium, elux of potassium, and reduction in membrane resting potential. As shock progresses, the loss of cellular integrity and the breakdown in cellular homeostasis result in cellular death. These pathologic events give rise to a cascade of metabolic features including hyperkalemia, hyponatremia, azotemia, hyper- or hypoglycemia, and lactic acidosis.

Inflammation plays an important role in several dierent types of shock, especially in septic shock, but also in shock associated with anaphylaxis, burns, trauma, and cardiogenic causes.15 Previously, the systemic inflammatory response syndrome was part of the definition of sepsis, but that changed with a revised definition of sepsis, Sepsis-3, in 2016.16 See Chapter 151, “Sepsis,” for detailed discussion.

As global tissue hypoxia progresses, shock ensues, followed by the multiorgan dysfunction syndrome, which is manifested by renal failure, respiratory failure, myocardial depression, liver failure, and then disseminated intravascular coagulation. The fulminant progression from global tissue hypoxia to multiorgan dysfunction syndrome is determined by the severity of inadequate tissue perfusion and the balance of anti-inflammatory and proinflammatory mediators (Figure 12-1).

FIGURE 12-1. The pathophysiology of shock, the inflammatory response, and multiorgan dysfunction.

Loading [Contrib]/a11y/accessibility-menu.js CLINICAL FEATURES

HISTORY AND COMORBIDITIES

Although the clinical presentation of a patient in shock and the underlying cause may be quite apparent (e.g., acute myocardial infarction, anaphylaxis, or hemorrhage), it may be diicult to obtain a history from patients in shock. Assistance with medical history from EMS, family, or other sources may help determine the cause of shock. Some patients in shock may have few symptoms other than generalized weakness, lethargy, or altered mental status. If the patient is unresponsive, consider trauma as a primary or secondary complication.

PHYSICAL EXAMINATION

Shock is usually associated with systemic arterial hypotension—systolic blood pressure <90 mm Hg. Blood pressure may not drop if there is an increase in peripheral vascular resistance in the presence of decreased CO with inadequate tissue hypoperfusion. For this reason, blood pressure is an insensitive marker for global tissue hypoperfusion. Shock may occur with a normal blood pressure, and hypotension may occur without shock. No single vital sign is diagnostic of shock, and blood pressure is particularly insensitive in the presence of peripheral vascular disease, tachycardia with a small pulse pressure, or cardiac dysrhythmias. Composite physical findings are useful in the assessment of shock (Table 12-2).

Loading [Contrib]/a11y/accessibility-menu.js TABLE 12-2 Composite Physical Examination Findings in Shock

Temperature Hyperthermia or hypothermia may be present. Endogenous hypothermia (hypometabolic shock) must be distinguished from exogenous environmental hypothermia.

Heart rate Usually elevated; however, paradoxical bradycardia can be seen in shock states due to hypoglycemia, β-blocker use, and preexisting cardiac disease.

Systolic blood May actually increase slightly when cardiac contractility increases in early shock and then fall as shock advances. pressure

Diastolic blood Correlates with arteriolar vasoconstriction and may rise early in shock and then fall when cardiovascular pressure compensation fails.

Pulse pressure Increases early in shock and decreases before systolic pressure begins to drop.

Mean arterial Oen low, <65 mm Hg. blood pressure

CNS Acute delirium, restlessness, disorientation, confusion, and coma secondary to a decrease in cerebral perfusion pressure.

Skin/capillary Pallor, pale, dusky, clammy, cyanosis, sweating, cool, and capillary refill time >2–3 s. refill

Cardiovascular Neck vein distention or flattening depending on the type of shock. Tachycardia and arrhythmias. An S3 may result from high-output states. Decreased coronary perfusion pressures can lead to ischemia, decreased ventricular compliance, increased le ventricular diastolic pressure, and pulmonary edema.

Respiratory Tachypnea, increased minute ventilation, increased dead space, bronchospasm, and hyper- or hypocapnia with progression to respiratory failure.

Splanchnic Ileus, GI , pancreatitis, and mesenteric ischemia can occur due to low-flow states. organs

Renal Reduced glomerular filtration rate. Renal blood flow redistributes from the renal cortex toward the renal medulla leading to oliguria.

Metabolic Lactic acidosis, hyperglycemia, hypoglycemia, and hyperkalemia. As shock progresses, metabolic acidosis occurs with concurrent respiratory compensation.

DIAGNOSIS

LABORATORY EVALUATION

No single laboratory value is sensitive or specific for shock. Laboratory studies are driven by the clinical presentation and the Loading [Contrib]/a11y/accessibility-menu.js presumptive cause. Common studies are listed in Table 12-3. Arterial blood gases are useful to assess acid-base status and ventilation and oxygenation concerns, whereas a venous blood gas is limited to acid-base information. A rise in serum lactate correlates with mortality in many shock states12; typically this is due to anaerobic metabolism, but nonhypoxic causes of lactic acidosis due to cellular dysfunction occur in shock states. A wide range of laboratory abnormalities may be encountered in shock. The incidence of adrenal dysfunction can be as high as 30% in this subset of patients. Most abnormal values merely point to the particular organ system that is contributing to, or being aected by, the shock state.

TABLE 12-3 Initial Diagnostic Studies to Evaluate a Patient in Shock*

CBC with dierential Electrolytes, glucose, calcium, magnesium, phosphorus BUN, creatinine Serum lactate ECG Urinalysis Chest radiograph Coagulation studies: prothrombin time, PTT, INR Arterial blood gas (pH, carbon dioxide and oxygen levels, base deficit) Hepatic function panel Cultures: blood, urine, suspicious wounds, quantitative sputum culture Cortisol level Pregnancy test CT of chest/abdomen/pelvis as indicated by history, physical exam

*Ordering of tests should be individualized by patient presentation and history (see also Chapter 151, “Sepsis”).

IMAGING

Chest Radiograph

The portable anteroposterior view chest radiograph is oen used in the evaluation of unstable patients to avoid transporting the patient during resuscitation. While limitations exist, evaluation of the heart size, presence of pulmonary edema, free air under the diaphragm, pneumothorax, infiltrates, or eusions may provide useful clinical information.

US

Bedside US assessment is an important tool for developing a dierential diagnosis, assessing volume status, defining cardiac function, and assisting with procedures. Various US methods are described to determine overall volume status by assessing right- sided filling pressures, including measuring inferior vena cava respiratory variation or end-expiratory vena cava respiratory variation, and other methods.

Bedside cardiac US to assess le ventricular ejection fraction can assist with determining the cause of shock. Emergency physicians trained in focused bedside cardiac US can provide an estimated ejection fraction with high relative correlation to cardiologists.17

US may also be used to assess for vascular emergencies. Identifying an abdominal aortic aneurysm on US may lead to further evaluation. Findings of a deep vein thrombosis may increase the suspicion for a pulmonary embolism.

Loading [Contrib]/a11y/accessibility-menu.js Additional US protocols, such as the Abdominal and Cardiac Evaluation with Sonography in Shock protocol and the Rapid Ultrasound in Shock protocol, have been formulated using many of the aforementioned concepts.17-20 The Abdominal and Cardiac Evaluation with Sonography in Shock protocol looks at cardiac function, inferior vena cava dynamics, pulmonary congestion, lung sliding and consolidation, abdominal free fluid, abdominal aortic aneurysm, and leg venous thrombosis to assist in dierential diagnosis generation or narrowing.17,18 The Rapid Ultrasound in Shock exam involves a three-part bedside physiologic assessment simplified as the pump (cardiac), the tank (volume status), and the pipes (arterial and venous).19-21 However, as with any US intervention, operator competency is essential.

CT

Although CT is an accurate and noninvasive approach for detecting internal pathology, patients must travel from the ED to the radiology suite, which may be unadvisable in unstable shock. The potential benefits of CT must be weighed against the associated risks, including concerns about renal function due to and contrast-induced nephropathy. CT scans without IV contrast will add some information to the clinical picture, although not to the degree of a scan with IV contrast.

HEMODYNAMIC MONITORING

Hemodynamic monitoring helps assess the severity of shock and the response to treatment. Monitoring capabilities should initially include pulse oximetry, ECG monitoring, and noninvasive blood pressure monitoring. In the critical care arena, intra-arterial blood pressure monitoring, end-tidal carbon dioxide monitoring, central venous pressure, and central venous oxygen saturation from the superior vena cava (ScvO2) monitoring are frequently used. When obtaining central access, the average access time, number of attempts, and mechanical complications are reduced when a US-assisted approach is used.22

TREATMENT

See Chapter 151, “Sepsis,” for treatment of sepsis. Comprehensive and timely ED care can significantly decrease the predicted mortality of critically ill patients in as little as 6 hours of treatment.23 Application of an algorithmic approach to optimize hemodynamic end points with early goal-directed therapy in the ED reduced mortality by 16% in patients with severe sepsis or septic shock in 2001.24 That original study, the Surviving Sepsis Campaign that followed,25 and other algorithmic eorts26 have changed the approach to sepsis and shock care on a worldwide basis. Two large, multicenter, randomized controlled trials published in 2014 failed to show additional benefits to a rigid algorithmic approach.27,28 However, we attempt to present below the most beneficial aspects of shock care demonstrated by the medical progress of the past 15 years. The ABCDE tenets of shock resuscitation are establishing airway, controlling the work of breathing, optimizing the circulation, ensuring adequate oxygen delivery, and achieving end points of resuscitation.29

ESTABLISHING THE AIRWAY

Airway control is best obtained through endotracheal intubation. Sedatives used to facilitate intubation may cause arterial vasodilatation, venodilation, or myocardial suppression and may result in hypotension. Positive-pressure ventilation reduces preload and CO. The combination of sedative agents and positive-pressure ventilation will oen lead to hemodynamic collapse. To avoid this unwanted situation, initiate volume resuscitation and vasoactive agents before intubation and positive-pressure ventilation.

CONTROLLING THE WORK OF BREATHING

Control of breathing is required when significant work of breathing accompanies shock. Respiratory muscles are significant consumers of oxygen during shock and contribute to lactate production. Mechanical ventilation and sedation allow for adequate oxygenation, improvement of hypercapnia, and assisted, controlled, synchronized ventilation. All of these treatments decrease the Loading [Contrib]/a11y/accessibility-menu.js work of breathing and improve survival. When starting mechanical ventilation on a patient, it is essential to consider the patient’s compensatory minute ventilation prior to intubation to ensure appropriate initial settings are selected. Aer a patient is placed on mechanical ventilation, obtain an arterial blood gas to evaluate acid-base status, oxygenation, and ventilation. Neuromuscular blocking agents should be considered to further decrease respiratory muscle oxygen consumption and preserve oxygen delivery to vital organs, especially if patients are severely hypoxemic due to acute respiratory distress syndrome.30

OPTIMIZING THE CIRCULATION

Fluids

Circulatory or hemodynamic stabilization begins with intravascular access through large-bore peripheral venous lines. The Trendelenburg position does not improve cardiopulmonary performance compared with the supine position. It may worsen pulmonary gas exchange and predispose to aspiration. Passive leg raising above the level of the heart with the patient supine may be eective. If passive leg raising results in an increase in blood pressure or CO, fluid resuscitation is indicated.31

Fluid resuscitation should begin with isotonic crystalloid.32 Balanced crystalloids, such as lactated Ringer’s solution, may oer a small mortality advantage over normal saline solution (14.3% vs. 15.4% in a large trial of 15,802 critically ill patients).33 The amount and rate of infusion are determined by an estimate of the hemodynamic abnormality. Most patients in shock have either an absolute or relative volume deficit. The exception is the patient in cardiogenic shock with pulmonary edema. Administer fluid rapidly (over 5 to 20 minutes), in set quantities of 500 or 1000 mL of normal saline, and reassess the patient aer each bolus. Patients with a modest degree of hypovolemia usually require an initial 20 to 30 mL/kg of isotonic crystalloid, and current Centers for Medicare and Medicaid Services sepsis guidelines require 30 mL/kg; however, there are few data to support this uniform recommendation.34 More fluids may be needed for profound volume deficits. It is common for patients in septic shock to receive 6 L of crystalloid in the first 24 hours of hospital care. For large fluid volumes, consider using lactated Ringer’s or Plasma-Lyte® to avoid hyperchloremic metabolic acidosis associated with 0.9% sodium chloride solution35,36 (see later section “Controversies of Treatment”). In clinical situations where hypochloremia can be predicted, such as from GI losses due to vomiting or from urinary excretion due to diuretics, there may be an advantage to 0.9% sodium chloride rehydration.

Central venous access may aid in assessing volume status (preload) and monitoring ScvO2. It is also the preferred route for the long- term administration of certain vasopressor therapy. However, there is no need for universal central access in patients with septic shock, and the need for central access should be individually determined.26

Vasopressors

Vasopressors are used when there has been an inadequate response to volume resuscitation or if there are contraindications to volume infusion.33 Vasopressors are most eective when the vascular space is “full” and least eective when the vascular space is depleted. Patients with chronic hypertension may be at greater risk of renal injury at lower blood pressures; however, in others, there appears to be no mortality benefit in raising mean arterial pressure above the 65 to 70 mm Hg range.37,38

Vasopressor agents have variable eects on the α-adrenergic, β-adrenergic, vasopressin, and dopaminergic receptors (Table 12-4). Although vasopressors improve perfusion pressure in the large vessels, they may decrease capillary blood flow in certain tissue beds, especially the GI tract and peripheral vasculature. If multiple vasopressors are used, they should be simplified as soon as the best therapeutic agent is identified. In addition to a vasopressor, an inotrope may be needed to directly increase CO by increasing contractility and stroke volume. Careful vasopressor infusion initially through a peripheral IV is unlikely to result in tissue injury and will improve the time to achieve hemodynamic stability.39,40

Loading [Contrib]/a11y/accessibility-menu.js TABLE 12-4 Commonly Used Vasoactive Agents (all vasopressors increase myocardial oxygen demand; most should be titrated to desired eect)

Cardiac Cardiac Drug Dose Action Vasoconstriction Vasodilation Contractility Output

Dobutamine 2.0–20.0 β1, some β2 ++++ + ++ Increases micrograms/kg/min and α1 in large dosages

Side eects and Inotrope only; causes tachydysrhythmias, GI distress, hypotension in volume-depleted patients; less peripheral comments vasoconstriction than dopamine; fewer arrhythmias than isoproterenol

Dopamine 0.5–20 α, β, and ++ at 2.5–5 ++ at 5–20 + at 0.5–2.0 Usually micrograms/kg/min dopaminergic micrograms/kg/min micrograms/kg/min micrograms/kg/min increases

Side eects and Tachydysrhythmias; a cerebral, mesenteric, coronary, and renal vasodilator at low doses; Surviving Sepsis Campaign comments second line, lot of overlap with α/β/dopaminergic receptors and dose

Epinephrine 2–10 α and β ++++ at 0.5–8 ++++ at >8 +++ Increases micrograms/min micrograms/kg/min micrograms/kg/min

Side eects and Causes tachydysrhythmia, leukocytosis; increases myocardial oxygen consumption; may increase lactate; no real comments maximum dose

Isoproterenol 0.01–0.1 β1 and some ++++ 0 ++++ Increases microgram/kg/min β2

Side eects and Inotrope; causes tachydysrhythmia, facial flushing, hypotension in hypovolemic patients; increases myocardial oxygen comments consumption

Norepinephrine 0.5–50 Primarily α1, ++ ++++ 0 Slightly micrograms/min increases some β1

Side eects and Useful when loss of venous tone predominates; first-line agent for most situations comments

Phenylephrine 10–200 Pure α 0 ++++ 0 Decreases micrograms/min

Side eects and Reflex bradycardia, headache, restlessness, excitability, rarely arrhythmias; used on patients in shock with tachycardia or comments supraventricular arrhythmias; not good comparatively for septic shock

Vasopressin 0.01–0.04 units/min Directly 0 ++++ 0 0

stimulates V1 receptor on smooth

Loading [Contrib]/a11y/accessibility-menu.js muscle Cardiac Cardiac Drug Dose Action Vasoconstriction Vasodilation Contractility Output

Side eects and Primarily vasoconstriction; usually started at max dose and not titrated, typically added to norepinephrine comments

Note: 0 = no eect; + = mild eect; ++ = moderate eect; +++ = marked eect; ++++ = very marked eect.

ENSURING ADEQUATE OXYGEN DELIVERY

Control of oxygen consumption is important in restoring the balance of oxygen supply and demand to the tissue (oxygen consumption equation). A hyperadrenergic state results from the compensatory response to shock, physiologic stress, pain, cold treatment rooms, and anxiety. Pain further suppresses myocardial function, impairing oxygen delivery and increasing consumption. Providing analgesia, muscle relaxation, warm covering, anxiolytics, and even paralytic agents, when appropriate, decreases this inappropriate systemic oxygen consumption.

Once blood pressure is stabilized through optimization of preload and aerload, oxygen delivery can be assessed and further manipulated. Restore arterial oxygen saturation to ≥91%. In shock states, consider a transfusion of packed red blood cells to maintain hemoglobin ≥7 grams/dL.34 If CO can be assessed, it should be increased using volume infusion or inotropic agents in incremental amounts until venous oxygen saturation (mixed venous oxygen saturation [SvO2] or ScvO2) and lactate are normalized.

Sequential examination of lactate and SvO2 or ScvO2 is a method to assess adequacy of a patient’s resuscitation. Continuous measurement of SvO2 or ScvO2 can be used in the ED, although results from the ProCESS trial question the need for this in resuscitation management.28 A variety of technologic tools may be used to assess tissue perfusion during resuscitation.41-47 These technologies may be available in some EDs, but are more oen found in intensive care units. Transfer of the patient to the intensive care unit should not be delayed so that monitoring devices can be placed in the ED.23,47

END POINTS OF RESUSCITATION

The goal of resuscitation is to use hemodynamic and physiologic values to guide therapy in order to maximize survival and minimize morbidity. No therapeutic end point is universally eective, and only a few have been tested in prospective trials, with mixed results.28,37,38 Hypotension at ED presentation is associated with poor outcomes.48 Noninvasive parameters, such as blood pressure, heart rate, and urine output, may underestimate the degree of remaining hypoperfusion and oxygen debt, so the use of additional physiologic end points may be informative.28,37,38,48 A goal-directed approach of mean arterial pressure >65 mm Hg, central venous pressure of 8 to 12 mm Hg, ScvO2 >70%, and urine output >0.5 mL/kg/h during ED resuscitation of septic shock has been shown to decrease mortality, but which of the metrics accounts for the mortality decrease remains in question.24,27,28,34 Source control, whether with infection, hemorrhage, or other state of shock, is essential in the initial stages of management. If shock or hypotension persists, reassessment at the patient’s bedside is essential while considering the important issues in Table 12- 5.

Loading [Contrib]/a11y/accessibility-menu.js TABLE 12-5 Questions to Answer If There Is Persistent Shock or Hypotension

Equipment and monitoring Is the patient appropriately monitored? Is there an equipment malfunction, such as dampening of the arterial line or disconnection from the transducer? Is the IV tubing into which the vasopressors are running connected appropriately? Are the vasopressor infusion pumps working? Are the vasopressors mixed adequately and in the correct dose?

Patient assessment Do mentation and clinical appearance match the degree of hypotension? Is the patient adequately volume resuscitated? Does the patient have a pneumothorax aer placement of central venous access? Has the patient been adequately assessed for an occult penetrating injury? Is there hidden bleeding from a ruptured spleen, large-vessel aneurysm, or ectopic pregnancy? Does the patient have adrenal insuiciency? Is the patient allergic to the medication just given or taken before arrival? Is there cardiac tamponade in the dialysis patient or cancer patient? Is there associated acute myocardial infarction, aortic dissection, or pulmonary embolus?

CONTROVERSIES OF TREATMENT

Fluid Therapy

Rapid restoration of fluid deficits modulates inflammation and, if the condition progresses to shock, decreases the need for subsequent vasopressor therapy, steroid administration, and invasive monitoring (e.g., pulmonary artery catheterization and arterial line placement).33,46 Although there is general agreement that volume therapy is an integral component of early resuscitation, there is a lack of consensus for the type of fluid, standards of volume assessment, and end points despite care mandates that dictate otherwise.49 Table 12-6 compares the most commonly used fluid therapies.50-54

Loading [Contrib]/a11y/accessibility-menu.js TABLE 12-6 Fluid Therapy

Crystalloids

Normal Slightly hyperosmolar containing 154 mEq/L of both sodium and chloride. saline (NS) Risk of inducing hyperchloremic metabolic acidosis when given in large amounts due to relatively high chloride concentration.

Lactated Sodium 130 mEq/L, potassium 4 mEq/L, calcium 3 mEq/L, chloride 109 mEq/L, lactate 28 mEq/L. Lactate can accept a Ringer’s (LR) proton and subsequently be metabolized to carbon dioxide and water by the liver, leading to release of carbon dioxide in the lungs and excretion of water by the kidneys. LR results in a buering of the acidemia that is advantageous over NS. Theoretical risk of inducing hyperkalemia in patients with renal insuiciency or renal failure due to small potassium content (very small amount).

Plasma-Lyte® Balanced pH 7.4, sodium 140 mEq/L, chloride 98 mEq/L, potassium 5 mEq/L, magnesium 3 mEq/L, gluconate 23 mEq/L, acetate 27 mEq/L.

Colloids

Albumin Derived from human plasma. Available in varying strengths from 4% to 25%. Multiple studies have shown that there is no outcome dierence whether colloids or crystalloids are used. Colloids cost significantly more than crystalloids. One study actually showed an increase in mortality in trauma patients with head injury.

Hydroxyethyl Synthetic colloid derived from hydrolyzed amylopectin. starch These agents should be avoided in sepsis. Many harmful eects: renal impairment at recommended doses and impairment of long-term survival at high doses, coagulopathy and bleeding complications from reduced factor VIII and von Willebrand factor levels, impaired platelet function.

Colloids are high-molecular-weight solutions that increase plasma oncotic pressure. Colloids can be classified as either natural (albumin) or artificial (starches, dextrans, and gelatins). Due to their higher molecular weight, colloids stay in the intravascular space significantly longer than crystalloids. The intravascular half-life of albumin is 16 hours versus 30 to 60 minutes for normal saline and lactated Ringer’s solution.32,50-54

Resuscitation with crystalloids requires two to four times more volume than colloids.32,49,50 The outcome advantage between crystalloid and colloids continues to remain unresolved in sepsis, despite multiple studies.32,49-52 Due to the equivalency and the higher cost of colloids, crystalloids would seem to be a better choice for resuscitation in the ED.

Crystalloid type aects acid-base status and may have an eect on mortality, acute kidney injury, and the need for renal replacement therapy.35,36,55-58 Two large, single-center, randomized controlled trials compared normal saline to balanced salt solutions (lactated Ringer’s or Plasma-Lyte®) in the resuscitation of patients in the ED prior to admission and found fewer major adverse events (persistent renal dysfunction, new renal replacement therapy, or mortality) in the balanced salt group.55,56 The dierence favoring fewer adverse eects in the balance salt group was small (10.3% vs. 11.1% in those admitted to the intensive care unit and 4.7% vs. 5.6% in those admitted to the floor).55,56 Since then, a meta-analysis including those two studies, plus four Loading [Contrib]/a11y/accessibility-menu.js prior randomized controlled trials, totaling 19,332 patients, found no significant dierences in the outcome measures of intensive care unit mortality, acute kidney injury, and new renal replacement therapy when comparing 0.9% normal saline to balanced salt solutions.59 Several ongoing randomized controlled trials comparing IV resuscitation fluids are pending completion.57,60,61 A randomized controlled trial of 157 patients receiving IV hydration with 2 L of 0.9% normal saline versus lactated Ringer’s solution before presumed discharge from ED found no dierence in outcomes.62 The majority of the patients were healthy, with hypovolemia from vomiting or diarrhea.62 Most consistent in all the cited trials is a hyperchloremic acidosis associated with large- volume resuscitation with normal saline and, as of yet, no measured adverse eects of resuscitation with balanced salt solutions, despite a theoretical potential for metabolic alkalosis, hypotonicity from fluids with lactate, and cardiotoxicity from fluids with acetate.63

Sodium Bicarbonate

Bicarbonate administration shis the oxygen-hemoglobin dissociation curve to the le, impairs tissue unloading of hemoglobin- bound oxygen, and may worsen intracellular acidosis. However, despite no definitive clinical trials supporting benefit but perhaps harm, many clinicians remain uncomfortable withholding bicarbonate if the pH is <7.00.64,65 Animal studies of profound acidosis demonstrate decreased ventricular contractility and systolic blood pressure.13 If bicarbonate is given, recognize the risk of paradoxical intracerebral intracellular acidosis in the process. Consider situations, such as end-stage renal disease and renal tubular acidosis, that cannot reclaim bicarbonate through normal renal processes and whether bicarbonate may be indicated.

DISPOSITION

TRANSITION TO THE INTENSIVE CARE UNIT

Early recognition, treatment, and subsequent transfer of critically ill patients to the intensive care unit improves patient outcomes and improves ED throughput.23,47 Communicate and document all ED resuscitative eorts to the critical care team. Even when resuscitation is systematic and thoughtful, miscommunication can undo the benefits of initial ED treatment. Ideally, before transfer, verbally communicate and document a system-oriented problem list with an assessment and plan, including all procedures and complications. For prolonged or “boarded” ED stays, constantly reassess the critically ill patient, ensure that care plans are continuing, and consider creating a critical care patient checklist. Oen, this will entail ordering tests that are not commonly performed on ED patients or ordering subsequent doses of medicines, particularly antibiotics.

PROGNOSIS

Some clinical variables are associated with poor outcome, such as severity of shock, temporal duration, underlying cause, preexisting vital organ dysfunction, and reversibility. Early recognition, intervention, source control, and smooth care transitions optimize outcomes. While associated morbidity and mortality remain high for patients with shock, integration of protocol-based care pathways with ongoing refinement in response to new information may lead to continued reductions over time.23,28,34,47 Additional outcome predictions related to physiologic scoring systems, ED-based shock interventions, and the balance between invasive and noninvasive or minimally invasive strategies are still being studied.23,66,67

REFERENCES

1. Holler JG, Bech CN, Henriksen DP, Mikkelsen S, Pedersen C, Lassen AT: Nontraumatic hypotension and shock in the emergency department and the prehospital setting, prevalence, etiology, and mortality: a systematic review. PLoS One 10: e0119331, 2015. [PubMed: 25789927]

2. Fleischmann C, Scherag A, Adhikari JKJ, et al: Assessment of global incidence and mortality of hospital-treated sepsis. Current estimates and limitations. Am J Resp Crit Care Med 193: 259, 2016. [PLuobadMinegd [:C 2o6nt4r1ib4]/2a9112y]/ a ccessibility-menu.js 3. Jeger RV, Radovanovic D, Hunziker PR, et al: Ten-year trends in the incidence of cardiogenic shock. Ann Intern Med 149: 618, 2008. [PubMed: 18981487]

4. Thiele H, Zeymer U, Neumann FJ, et al: Intraaortic balloon support for myocardial infarction with cardiogenic shock. N Engl J Med 367: 1287, 2012. [PubMed: 22920912]

5. Guly HR, Bouamra O, Lecky FE. The incidence of neurogenic shock in patients with isolated spinal cord injury in the emergency department. Resuscitation 76:57, 2008. [PubMed: 17688997]

6. Russell JA, Rush B, Boyd J: Pathophysiology of septic shock. Crit Care Clin 34: 43, 2018. [PubMed: 29149941]

7. Biering-Sørensen F, Biering-Sørensen T, Liu N, et al: Alterations in cardiac autonomic control in spinal cord injury. Auton Neurosci 209: 4, 2018. [PubMed: 28228335]

8. Ashruf JF, Bruining HA, Ince C: New insights into the pathophysiology of cardiogenic shock: the role of the microcirculation. Curr Opin Crit Care 19: 381, 2013. [PubMed: 23995131]

9. Holler JG, Jensen HK, Henriksen DP, et al: Etiology of shock in the emergency department; a 12 year population based cohort study. Shock. 51:60, 2019. [PubMed: 27984523]

10. Kheng CP, Rahman NH: The use of end-tidal carbon dioxide monitoring in patients with hypotension in the emergency department. Int J Emerg Med 5: 31, 2012. [PubMed: 22828152]

11. Holt BD, Walsh RA: Normal physiology of the cardiovascular system, in Fuster V, Walsh RA, Harrington RA (eds): Hurst’s the Heart, 13th ed. New York, NY: McGraw Hill; 2011.

12. Shapiro NI, Howell MD, Talmor D, et al: Serum lactate as a predictor of mortality in emergency department patients with infection. Ann Emerg Med 45: 524, 2005. [PubMed: 15855951]

13. Bonanno FG: Hemorrhagic shock: the “physiology approach.” J Emerg Trauma Shock 5: 285, 2012. [PubMed: 23248495]

14. Oedorf K, Day D, Lior Y, et al: Serum lactate predicts adverse outcomes in emergency department patients with and without infection. West J Emerg Med 18: 258, 2017. [PubMed: 28210362]

15. Parenica J, Jarkovsky J, Malaska J, et al: Infectious complications and immune/inflammatory response in cardiogenic shock: a prospective observational study. Shock 47: 165, 2017. [PubMed: 27749762]

Loading [Contrib]/a11y/accessibility-menu.js 16. Singer M, Deutschman CS, Seymour CW, et al: The third international consensus definitions for sepsis and septic shock (Sepsis-3). JAMA 315: 801, 2016. [PubMed: 26903338]

17. Arntfield RT, Millington SJ: Point of care cardiac ultrasound applications in the emergency department and intensive care unit: a review. Curr Cardiol Rev 8: 98, 2012. [PubMed: 22894759]

18. Atkinson PRT, McAuley DJ, Kendall RJ, et al: Abdominal and Cardiac Evaluation with Sonography in Shock (ACES): an approach by emergency physicians for the use of ultrasound in patients with undierentiated hypotension. Emerg Med J 26: 87, 2009. [PubMed: 19164614]

19. Seif D, Perera P, Mailhot T, et al: Bedside ultrasound in resuscitation and the rapid ultrasound in shock protocol. Crit Care Res Pract 2012: 503254, 2012. [PubMed: 23133747]

20. Mok KL: Make it SIMPLE: enhanced shock management by focused cardiac ultrasound. J Intensive Care 4: 51, 2016. [PubMed: 27529027]

21. Birkhahn RH, Gaeta TJ, Terry D, et al: Shock index in diagnosing early acute hypovolemia. Am J Emerg Med 23: 323, 2005. [PubMed: 15915406]

22. Karakitsos D, Labropoulos N, DeGroot E, et al: Real-time ultrasound-guided catheterization of the internal jugular vein: a prospective comparison with the landmark technique in critical care patients. Crit Care 10: R162, 2006. [PubMed: 17112371]

23. Institute of Medicine: IOM report: the future of emergency care in the United States health system. Acad Emerg Med 13: 1081, 2006. [PubMed: 17015421]

24. Rivers E, Nguyen B, Havstad S, et al: Early goal-directed therapy in the treatment of severe sepsis and septic shock. N Engl J Med 345: 1368, 2001. [PubMed: 11794169]

25. Dellinger RP, Levy MM, Carlet JM, et al: Surviving Sepsis Campaign: international guidelines for management of severe sepsis and septic shock: 2008. Crit Care Med 36: 296, 2008. [PubMed: 18158437]

26. Sebat F, Musthafa AA, Johnson D, et al: Eect of a rapid response system for patients in shock on time to treatment and mortality during 5 years. Crit Care Med 35: 2568, 2007. [PubMed: 17901831]

27. ARISE Investigators, ANZICS Clinical Trials Group: Goal-directed resuscitation for patients with early septic shock. N Engl J Med 371: 1496, 2014. [PubMed: 25272316]

28. The ProCESS Investigators: A randomized trial of protocol-based care for early septic shock. N Engl J Med 370: 1683, 2014. [PubMed: 24635773]

Loading [Contrib]/a11y/accessibility-menu.js 29. Balas MC, Vasilevskis EE, Burke WJ, et al: Critical care nurses’ role in implementing the “ABCDE bundle” into practice. Crit Care Nurse 32: 35, 2012. [PubMed: 22467611]

30. Foral JM, Voillet F, Pulina D, et al: Ventilator-associated pneumonia and ICU mortality in severe ARDS patients ventilated according to a lung-protective strategy. Crit Care 16: R65, 2012. [PubMed: 22524447]

31. Monnet X, Rienzo M, Osman D, et al: Passive leg raising predicts fluid responsiveness in critically ill patients. Crit Care Med 34: 1402, 2006. [PubMed: 16540963]

32. Annane D, Siami S, Jaber S, et al: Eects of fluid resuscitation with colloids vs crystalloids on mortality in critically ill patients presenting with hypovolemic shock: the CRISTAL randomized trial. JAMA 310: 1809, 2013. [PubMed: 24108515]

33. Semler MW, Self WH, Wanderer JP, et al: Balanced crystalloids versus saline in critically ill adults. N Engl J Med 378: 829, 2018. [PubMed: 29485925]

34. Dellinger RP, Levy MM, Rhodes A, et al: Surviving Sepsis Campaign: international guidelines for management of severe sepsis and septic shock, 2012. Intensive Care Med 39: 165, 2013. [PubMed: 23361625]

35. Yunos NM, Bellomo R, Hegarty C, Story D, Ho L, Bailey M: Association between a chloride-liberal vs chloride-restrictive intravenous fluid administration strategy and kidney injury in critically ill adults. JAMA 308: 1566, 2012. [PubMed: 23073953]

36. Raghunathan K, Bonavia A, Nathanson BH, et al: Association between initial fluid choice and subsequent in-hospital mortality during the resuscitation of adults with septic shock. Anesthesiology 123: 1385, 2015. [PubMed: 26414499]

37. Asfar P, Meziani F, Hamel JF, et al: High versus low blood-pressure target in patients with septic shock. N Engl J Med 370: 1583, 2014. [PubMed: 24635770]

38. Russell JA: Is there a good MAP for septic shock? N Engl J Med 370: 1649, 2014. [PubMed: 24635771]

39. Cardenas-Garcia J, Schaub KF, Belchikov YG, et al: Safety of peripheral intravenous administration of vasoactive medication. J Hosp Med 10: 581, 2015. [PubMed: 26014852]

40. Loubani OM, Green RS: A systematic review of extravasation and local tissue injury from administration of vasopressors through peripheral intravenous catheters and central venous catheters. J Crit Care 30: 653.e9, 2015. [PubMed: 25669592]

41. Cuschieri J, Rivers EP, Donnino MW, et al: Central venous-arterial carbon dioxide dierence as an indicator of cardiac index. Intensive Care Med 31: 818, 2005. [PubMed: 15803301] Loading [Contrib]/a11y/accessibility-menu.js 42. Weil MH, Nakagawa Y, Tang W, et al: Sublingual capnometry: a new noninvasive measurement for diagnosis and quantitation of severity of circulatory shock. Crit Care Med 27: 1225, 1999. [PubMed: 10446813]

43. Marik PE: Regional carbon dioxide monitoring to assess the adequacy of tissue perfusion. Curr Opin Crit Care 11: 245, 2005. [PubMed: 15928474]

44. Skarda DE, Mulier KE, Myers DE, et al: Dynamic near-infrared spectroscopy measurements in patients with severe sepsis. Shock 27: 348, 2007. [PubMed: 17414414]

45. Verdant C, De Backer D: How monitoring of the microcirculation may help us at the bedside. Curr Opin Crit Care 11: 240, 2005. [PubMed: 15928473]

46. Pinsky MR, Payen D: Functional hemodynamic monitoring. Crit Care 9: 566, 2005. [PubMed: 16356240]

47. Singer AJ, Thode HC Jr, Viccellio P, Pines JM: The association between length of emergency department boarding and mortality. Acad Emerg Med 18: 1324, 2011. [PubMed: 22168198]

48. Jones SE, Stiell IG, Nesbitt LP, et al: Nontraumatic out-of-hospital hypotension predicts in-hospital mortality. Ann Emerg Med 43: 106, 2004. [PubMed: 14707949]

49. Vincent JL, Gerlach H: Fluid resuscitation in severe sepsis and septic shock: an evidence-based review. Crit Care Med 32: S451, 2004. [PubMed: 15542955]

50. Caironi P, Tognoni G, Masson S, et al: Albumin replacement in patients with severe sepsis or septic shock. N Engl J Med 370: 1412, 2014. [PubMed: 24635772]

51. Finfer S, Bellomo R, Boyce N, et al: A comparison of albumin and saline for fluid resuscitation in the intensive care unit. N Engl J Med 350: 2247, 2004. [PubMed: 15163774]

52. American Thoracic Society: Evidence-based colloid use in the critically ill: American Thoracic Society Consensus Statement. Am J Respir Crit Care Med 170: 1247, 2004. [PubMed: 15563641]

53. Wiedermann CJ: Systematic review of randomized clinical trials on the use of hydroxyethyl starch for fluid management in sepsis. BMC Emerg Med 8: 1, 2008. [PubMed: 18218122]

54. Brunkhorst FM, Engel C, Bloos F, et al: Intensive insulin therapy and pentastarch resuscitation in severe sepsis. N Engl J Med 358: 125, 2008. [PubMed: 18184958]

Loading [Contrib]/a11y/accessibility-menu.js 55. Self WH, Semler MW, Wanderer JP, et al: Balanced crystalloids versus saline in noncritically ill adults. N Engl J Med 378: 819, 2018. [PubMed: 29485926]

56. Semler MW, Self WH, Wanderer JP, et al: Balanced crystalloids versus saline in critically ill adults. N Engl J Med 378: 829, 2018. [PubMed: 29485925]

57. Rochwerg B, Millen T, Austin P, et al: Fluids in Sepsis and Septic Shock (FISSH): protocol for a pilot randomised controlled trial. BMJ Open 7: e017602, 2017. [PubMed: 28729329]

58. Serpa Neto A, Martin Loeches I, Klanderman RB, et al: Balanced versus isotonic saline resuscitation-a systematic review and meta-analysis of randomized controlled trials in operation rooms and intensive care units. Ann Transl Med 5: 323, 2017. [PubMed: 28861420]

59. Zayed YZM, Aburahma AMY, Barbarawi MO, et al: Balanced crystalloids versus isotonic saline in critically ill patients: systematic review and meta-analysis. J Intensive Care 6: 51, 2018. [PubMed: 30140441]

60. McIntyre L, Taljaard M, McArdle T, et al: FLUID trial: a protocol for a hospital-wide open-label cluster crossover pragmatic comparative eectiveness randomized pilot trial. BMJ Open. 8: e022780, 2018. [PubMed: 30139908]

61. Zampieri FG, Azevedo LCP, Corrêa TD, et al: Study protocol for the Balanced Solution versus Saline in Intensive Care Study (BaSICS): a factorial randomised trial. Crit Care Resusc 19: 175, 2017. [PubMed: 28651514]

62. Friederich A, Martin N, Swanson MB, et al: Normal saline solution and lactated Ringer’s solution have a similar eect on quality of recovery: a randomized controlled trail. Ann Emerg Med 2018, August 23. [Epub ahead of print] [PubMed: 30146446]

63. Myburgh JA, Mythen MG: Resuscitation fluids. N Engl J Med 369: 1243, 2013. [PubMed: 24066745]

64. Kimmoun A, Novy E, Auchet T, et al: Hemodynamic consequences of severe lactic acidosis in shock states: from bench to bedside. Crit Care 19: 175, 2015. [PubMed: 25887061]

65. Kim HJ, Son YK, An WS: Eect of sodium bicarbonate administration on mortality in patients with lactic acidosis: a retrospective analysis. PLoS One 8: e65283, 2013. [PubMed: 23755210]

66. Rivers EP, Kruse JA, Jacobsen G, et al: The influence of early hemodynamic optimization on biomarker patterns of severe sepsis and septic shock. Crit Care Med 35: 2016, 2007. [PubMed: 17855815]

67. Nguyen HB, Banta JE, Cho TW, et al: Mortality predictions using current physiologic scoring systems in patients meeting criteria for early goal-directed therapy and the severe sepsis resuscitation bundle. Shock 30: 23, 2008. [PubMed: 18323748] Loading [Contrib]/a11y/accessibility-menu.js McGraw Hill Copyright © McGraw-Hill Education All rights reserved. Your IP address is 75.148.241.33 Terms of Use • Privacy Policy • Notice • Accessibility

Access Provided by: HCA Healthcare Silverchair

Loading [Contrib]/a11y/accessibility-menu.js